You are on page 1of 6

A Backstepping Design for Flight Path Angle Control

Ola Härkegård and S. Torkel Glad


Division of Automatic Control, Department of Electrical Engineering,
Linköpings universitet, SE-581 83 Linköping, Sweden.
E-mail: ola@isy.liu.se, torkel@isy.liu.se

Abstract all nonlinearities means that the resulting control law


may be much simpler than if feedback linearization had
A nonlinear approach to flight path angle control is been used.
presented. Using backstepping, a globally stabilizing
control law is derived. Although the nonlinear nature In this paper, the application is flight path angle con-
of the lift force is considered, the pitching moment to be trol. Using inherent characteristics of the lift force, we
produced is only linear in the measured states. Thus, will show that despite its nonlinear behavior around
the resulting control law is much simpler than if feed- the stall angle, it suffices to produce a pitching mo-
back linearization had been used. The free parameters ment that is linear in the measured states to stabilize
that spring from the backstepping design are used to the aircraft around the desired trajectory, regardless of
achieve a desired linear behavior around the operating the initial state.
point.
The remainder of the paper is organized as follows. Sec-
tion 2 presents backstepping in an informal setting. In
Section 3, a nonlinear model describing the longitudi-
1 Introduction nal aircraft dynamics in pure-feedback form, crucial to
the backstepping design, is derived. In Section 4, the
While many conventional flight control designs assume backstepping control law is derived in detail, and in
the aircraft dynamics to be linear about some nominal Section 5 the control law is evaluated using a realis-
flight condition, this paper deals with the problem of tic simulation model. Section 6 contains a comparison
controlling an aircraft explicitly considering its nonlin- with feedback linearization.
ear dynamics. Contributions along this line often use
feedback linearization [9], i.e., controllers based upon
dynamic inversion, to deal with the nonlinearities. The 2 Backstepping
control law is designed to cancel the nonlinearities so
that the closed loop system is rendered linear. For backstepping [4] to be applicable, the differential
equations describing the system dynamics need to have
Two motivating reasons for this can be found. First, a
a certain pure-feedback structure:
linear system is easy to analyze with respect to stabil-
ity and performance. Second, a linear system has the ẋ1 = f1 (x1 , x2 )
same dynamic properties independently of the operat- ..
ing point, e.g., speed, height, and angle of attack in the . (1)
aircraft case. The second is true only for linear systems ẋn−1 = fn−1 (x1 , . . . , xn−1 , xn )
and cannot be competed with. We will therefore focus
ẋn = fn (x1 , . . . , xn , u)
on the first property and present an alternative way of
finding control laws that guarantee stability and can be The reason for this is the way a backstepping control
tuned to achieve a certain closed loop behavior. law, u(x), is recursively constructed with the aim of
bringing the state vector, x, to the origin.
Our main tool will be backstepping [4], a Lyapunov
based design method that has received a lot of atten- To begin with, x2 is considered a virtual control of the
tion during the last decade. Compared to feedback lin- scalar ẋ1 subsystem. A stabilizing function, φ1 (x1 ), is
earization, backstepping offers a more flexible way of determined such that ẋ1 = f1 (x1 , φ1 (x1 )) has the (sta-
dealing with nonlinearities. Using Lyapunov functions, ble) properties desired. However, x2 is not available for
their impact on the system can be analyzed so that sta- control. The key property of backstepping is that it of-
bilizing, and thus in a sense useful, nonlinearities may fers a constructive way of forwarding the unattainable
be kept while harmful nonlinearities can be cancelled or control demand x2 = φ1 (x1 ) to a new virtual control
dominated by the control signal. Not having to cancel law x3 = φ2 (x1 , x2 ). If this could be satisfied, x1 and
6
x 10
2

L θ
α V 1
replacements
γ 0

L (N)
M
−1

−2
δ
−3
−40 −20 0 20 40 60
α (deg)
Figure 1: Illustration of longitudinal aircraft entities.
Figure 2: Typical lift force vs angle of attack relationship.
x2 would be successfully brought to the origin. This
recursive procedure is repeated until the actual control pitching moment. This will affect the angle of attack
variable u is reached after n steps, whereby a stabilizing to which the lift force is strongly related. Note that
control law, u = φn−1 (x), has been constructed. this is exactly the same assumption which is needed
for dynamic inversion to be applied to aircraft control
Along with the control law, a Lyapunov function is con-
[6]. Expressing L and M in terms of their aerodynamic
structed, proving the stability of the closed loop system.
coefficients, we have that
L = q̄SCL (α)
(6)
3 Aircraft model M = q̄Sc̄Cm (α, q, δ)
where q̄ = 12 ρV 2 is the dynamic pressure, ρ is the air
In this paper only control about the longitudinal axis density, S is the wing platform area, and c̄ is the mean
is considered. With this restriction, the equations of aerodynamic chord.
motion describing the aircraft take the form [10]

1
V̇ = (−D + FT cos α − mg sin γ) (2) 4 Design for flight path angle control
m
1
γ̇ = (L + FT sin α − mg cos γ) (3) 4.1 Deriving the control law
mV
θ̇ = q (4) The system description consists of Equations (3)-(5),
where the flight path angle γ is to be controlled. Given
1
q̇ = (M + FT ZT P ) (5) the reference value, γref , the angle of attack at steady
Iy state, α0 , is defined through γ̇ = 0, see Equation (3).
In the actual implementation, α0 is computed on-line
The state variables are V = airspeed, γ = flight path
and updated at each time step, using current values of
angle, θ = pitch angle, and q = pitch rate. α = θ − γ is
q̄, δ 1 , and FT . Although this pragmatic approach to de-
the angle of attack, FT is the engine thrust force, which
termining α0 is not guaranteed to converge, successful
contributes to the pitching moment due to a thrust
simulations provide an alibi.
point offset, ZT P . The aerodynamical effects on the
aircraft are captured by the lift and drag forces, L and In Equation (3), the γ dependence of the gravitational
D, and the pitching moment, M . See Figure 1, where term plays an insignificant role. Therefore we will only
δ represents the deflections of the control surfaces that consider its contribution at the equilibrium, and assume
are at our disposal. the dynamics to be
The controlled variables are V and γ. In the imple- 1
γ̇ = (L(α) + FT sin α − mg cos γref )
mented controller, evaluated in Section 5, airspeed con- mV (7)
trol is handled separately using results from [7]. Our , ϕ(α − α0 )
control design will therefore focus solely on Equations
(3)-(5). where ϕ(0) = 0. Using the sign properties of its com-
ponents, L(α), see Figure 2 for a typical example, and
To meet the structural demand in Section 2, we will FT sin α, we conclude that
neglect the dependence of the lift force on the pitch
αϕ(α) > 0, α 6= 0 (8)
rate and the control surface deflections. The principal
function of the control surfaces to produce an angular 1 In computing α0 , the lift force dependence on δ is included.
for all α of practical interest2 . This is a vital property an extra degree of freedom, where F is required to be
for the backstepping control design to follow which will positive definite and radially growing. This extension
be done in three steps. of the backstepping procedure is due to [5]. Hence,
c2 2 1 2
Step 1: First introduce the control error V2 = z + z + F (ξ), c2 > 0
2 1 2 2
z1 = γ − γref We compute its time derivative to find a new stabilizing
function, qdes .
whose dynamics are given by
V̇2 = c2 z1 ϕ(ξ) + z2 (q + c1 ϕ(ξ)) + F 0 (ξ)(−ϕ(ξ) + q)
ż1 = ϕ(α − α0 ) (9)
= (c2 z1 + c1 z2 − F 0 (ξ))ϕ(ξ) + (z2 + F 0 (ξ))q
considering a constant reference value. Now use the
control Lyapunov function (clf)
Although it may not be transparent, we can again find
1 a stabilizing function independent of ϕ. Choosing
V1 = z12
2
qdes = −c3 z2 , c3 > 0 (11)
to determine a stabilizing function, θdes , considering θ 0
F (ξ) = c4 ϕ(ξ), F (0) = 0, c4 > 0 (12)
as the control input of Equation (9).
where (12) is an implicit but perfectly valid choice of
V̇1 = z1 ϕ(α − α0 ) F , yields
= z1 ϕ(θ − z1 − γref − α0 )
= z1 ϕ(−(1 + c1 )z1 + θ + c1 z1 − γref − α0 ) V̇2 = (c2 z1 + (c1 − c3 c4 )z2 )ϕ(ξ) − c4 ϕ(ξ)2 − c3 z22
= z1 ϕ(−(1 + c1 )z1 ) < 0, z1 6= 0
Selecting
is achieved by selecting
c2 = −(1 + c1 )(c1 − c3 c4 ), c 3 c4 > c 1 (13)
θdes = −c1 z1 + γref + α0 , c1 > −1 (10)
gives the negative definiteness of V̇2 that we want since
The fact that c1 = 0 is a valid choice means that γ
feedback is not necessary for the sake of stabilization. V̇2 = (c1 − c3 c4 )ξϕ(ξ) − c4 ϕ(ξ)2 − c3 z22 < 0
However, it provides an extra degree of freedom for tun-
ing the closed loop performance. for all z1 , z2 6= 0, again using (8). The benefit of using
the extra term F (ξ) shows up in Equation (13). F (ξ) ≡
Step 2: Continue by introducing the deviation from 0 leads to c2 = −(1+c1 )c1 > 0 and the severe restriction
the virtual control law (10). c1 < 0 implying positive feedback from z1 .

z2 = θ − θdes = θ + c1 z1 − γref − α0 Step 3: The final backstepping iteration begins with


introducing the third residual
Defining
z3 = q − qdes = q + c3 z2
ξ = −(1 + c1 )z1 + z2
We also update the system description
we have that
ż1 = ϕ(ξ)
ż1 = ϕ(ξ)
ż2 = z3 − c3 z2 + c1 ϕ(ξ) (14)
ż2 = q + c1 ϕ(ξ)
ż3 = u + c3 (z3 − c3 z2 + c1 ϕ(ξ))
We will also need
where
ξ˙ = −ϕ(ξ) + q 1
u= (M (α, q, δ) + FT ZT P ) (15)
Iy
A regular backstepping design would proceed by ex- is regarded as the control variable. Also,
panding the control Lyapunov function with a term pe-
nalizing z2 . We do this, but also add a term F (ξ) as ξ˙ = −ϕ(ξ) + z3 − c3 z2
2 Strictly speaking, there exist limiting angles of attack, α
min
and αmax , beyond which either Equation (8) stops to hold or V3 is constructed by adding a term penalizing z3 .
where our simple aircraft model is no longer valid. Using the
1
Lyapunov function to be constructed the true region of attraction V3 = c5 V2 + z32 , c5 > 0
could, at least theoretically, be computed. 2
We get 4.2 Tuning the controller
2 The resulting control law (18) is parameterized by the
V̇3 = c5 (c1 − c3 c4 )ξϕ(ξ) − c4 ϕ(ξ) − c3 z22
 three parameters c1 , c3 , and c6 . It can be rewritten as
+ z3 (z2 + c4 ϕ(ξ))
+ z3 (u + c3 (z3 − c3 z2 + c1 ϕ(ξ))) u = −kx

≤ − c4 c5 ϕ2 (ξ) − c3 c5 z22 k = c1 c3 c6 c3 c6 c6 (20)
T

+ z3 (u + c3 z3 + (c5 − c23 )z2 + (c1 c3 + c4 c5 )ϕ(ξ)) x = γ − γref θ − γref − α0 q

Closed loop stability is ensured as long as the parameter


once again using (8). Select c5 = c23 to cancel the z2 z3 restrictions above are satisfied. What is then a suitable
cross-term and try yet another linear control law. choice?
u = −c6 z3 , c6 > c 3 (16) A natural course of action is to linearize the open loop
is a natural candidate and with this we investigate the system (3)-(5) around a proper operating point, and
resulting clf time derivative. determine a satisfactory linear control law, u = −kx.
Then solve (20) for c1 , c3 , and c6 and check if these are
V̇3 ≤ − c23 c4 ϕ2 (ξ) − c33 z22 − (c6 − c3 )z32
in agreement with the given restrictions. If so, the con-
+ (c1 c3 + c23 c4 )z3 ϕ(ξ) trol law (18) locally achieves the desired linear closed
The first three terms on the right hand side are all ben- loop behavior, and globally guarantees stability. If the
eficial. In order to investigate the impact of the last restrictions are violated, the guarantee for global sta-
cross-term, we complete the squares. bility is lost.
c1 c3 + c23 c4 This trial and error strategy for finding possible closed
V̇3 ≤ − c33 z22 − (c6 − c3 )(z3 − ϕ(ξ))2
2(c6 − c3 ) loop systems may seem unsatisfactory. However, the re-
(c1 c3 + c23 c4 )2 2 verse task of mapping the parameter restrictions above
− (c23 c4 − )ϕ (ξ) onto restrictions regarding, e.g., closed loop pole place-
4(c6 − c3 )
ments is not trivial and will not be dealt with here.
V̇3 is negative definite provided that the ϕ2 (ξ) coeffi-
cient is negative, which is true for 4.3 Realizing the controller
(c1 + c3 c4 )2 The derived control law (18) regards the angular pitch
c6 > c3 (1 + ) (17) acceleration, q̇, as the control variable, see Equation
4c3 c4
We now pick c4 to minimize this lower limit under the (15). In this section we deal with the issue of finding
constraints c4 > 0 and c3 c4 > c1 . control surface deflections, δ, that will produce the de-
sired pitching moment, M .
For c1 ≤ 0, we can make c1 + c3 c4 arbitrarily small
whereby Equation (17) reduces to Since exact knowledge of neither the pitching moment
nor the thrust force is available, we will include an extra
c6 > c 3 term to capture the residual. Thus we remodel Equa-
For c1 > 0 the optimal strategy can be shown to be tion (5) as
selecting c4 arbitrarily close to the bound c1 /c3 . This 1
yields q̇ = u = (q̄Sc̄Cm (α, q, δ) + FT ZT P ) + E (21)
Iy
c6 > c3 (1 + c1 )
where Cm and FT represent the information available
to us through an identified aircraft model. To capture
Summary: Let us summarize our results. The ini-
a constant bias, we model E as
tial system (3)-(5) was transformed into (14) through a
backstepping design. For the latter, the linear control Ė = 0 (22)
law (16) was shown to be globally stabilizing despite
the nonlinear nature of the system. In terms of the We note that the nonlinear first part of (21) (excluding
original state variables the control law becomes E) depends only on known quantities. This allows us
u = −c6 (q + c3 (θ + c1 (γ − γref ) − γref − α0 )) (18) to build an observer for (21)-(22) having linear error
dynamics [3], thus producing an estimate of E with an
c1 , c3 and c6 are user parameters restricted by exponentially decaying error.
c1 > −1
With E at our disposal, we solve (21) for Cm .
c3 > 0
 (19)
c3 c1 ≤ 0 Iy (u − E) − FT ZT P
c6 > Cm (α, q, δ) = (23)
c3 (1 + c1 ) c1 > 0 q̄Sc̄
The HIRM configuration offers two ways of controlling 30
the pitching coefficient; via the taileron at the back of
25
the aircraft, and via the canard wings. A number of

Flight path angle γ (deg)


ways of how to divide the work between these control 20
surfaces have been suggested. In this paper, however,
15
we put no effort into optimizing this aspect but sim-
ply ignore the canard wings, leaving everything to the 10
taileron. In the following, δ will therefore represent the
taileron’s angle of deflection, see Figure 1. 5
Reference trajectory
0 Designed linear response
The pitching coefficient, Cm , is usually a complicated Nonlinear HIRM simulation
function of the arguments involved. The standard way −5
0 5 10 15 20
of modeling it is to collect measurements for a number time (s)
of different situations and create a look-up table from 35
which intermediate values can be interpolated. Assum-
30
ing such a table is available to us, it can be used to

Angle of attack α (deg)


solve (23) for δ. Given measurements of α and q, the 25
left hand side typically becomes a close to monotonic
20
function of δ. This fits many numerical solvers well,
and the Van Wijngaarden-Dekker-Brent method [1, 2] 15
was chosen for solving (23). This method is a happy
marriage between bisection, providing sureness of con- 10

vergence, and inverse quadratic interpolation, providing 5


superlinear convergence in the best-case scenario.
0
0 5 10 15 20
Naturally, there are hard bounds on the possible time (s)
taileron deflections. For the HIRM, δ ∈ [−40◦ , 10◦ ] 10
is the possible range. In cases where (23) cannot be
satisfied, i.e., when enough pitching moment cannot be 0
Taileron deflection δ (deg)

produced, we saturate and choose the proper δ bound.


The effects of saturation have only been studied empir- −10
ically.
−20

5 Application −30

In this section we investigate the properties of the de- −40


0 5 10 15 20
rived control law, applied to High Incidence Research time (s)
Model (HIRM), a realistic model of a generic fighter air- 0.8
Angular pitch acceleration dq/dt (deg/s2)

craft. The model was originally released for the GAR- True dq/dt
0.6 Model excl. E
TEUR robust flight control challenge [8]. Model incl. E
0.4
For the simulations, a flight case where the aircraft is 0.2
in level flight at Mach 0.3 and at a height of 5000 ft was
0
chosen. Actuator and sensor dynamics, which were not
considered in the preceding design, were included in the −0.2
simulations. Also, a constant pitch coefficient error of
−0.4
−0.030 was added to Cm . The simulation results are
reproduced in Figure 3. −0.6

−0.8
The top picture displays the flight path angle time re- 0 5 10 15 20
time (s)
sponse to a staircase reference signal. Also plotted is
the time response of the desired linear closed loop sys-
tem, i.e., the control law applied to the linearization of Figure 3: From top to bottom: γ time response compared
(3)-(5). We see that for small changes in the reference, with the designed linear response, α time re-
the simulated response stays very close to the linear sponse, taileron deflection, and finally the true
one. For the large step in the reference, occurring af- q̇ compared with Equation (21), with and with-
ter 5 seconds, the simulated response has a longer rise out the model error estimate E.
time. This has two main causes. First, we see that the the control variable, we have shown a linear control law
control surfaces saturate at the beginning of the step, to be globally stabilizing despite the nonlinear nature
thus not producing the desired pitching moment. This of the lift force with respect to the angle of attack. The
consequently slows down the step response. Second, fact that the control law is linear makes it easy to tune
the angle of attack becomes very large during the ma- given a linearization of the open loop system. Using this
neuver. When α reaches its peak after 7 seconds, the the locally linear closed loop behavior can be selected.
true lift force value is significantly smaller than what
the linear model predicts, see Figure 2. In agreement Simulations using a realistic aircraft model including
with Equation (3), this also leads to slower γ dynamics. actuator and sensor dynamics, and a constant pitch
coefficient error proved the control law to be robust
In the bottom picture, the benefit of estimating the against the approximations made during the design.
model error E can be seen. The dotted curve, rep-
resenting the complete model of the angular pitch ac- A comparison with feedback linearization showed the
celeration introduced in Equation (21), initially has a potential benefits of using backstepping. The backstep-
bias error compared to the true q̇. This error is quickly ping control law is computationally much simpler, and
estimated whereafter the curve closely follows the solid is globally stabilizing while feedback linearization yields
curve, representing the true q̇. The latter was computed a singularity in the control law around the stall angle.
by differentiating q sensor data.

8 Acknowledgment
6 Backstepping vs feedback linearization
The authors would like to acknowledge the Defence
Having derived the globally stabilizing control law (18) Evaluation and Research Agency (DERA) of the United
for (3)-(5) using backstepping, it is rewarding to make Kingdom for supplying the HIRM model.
a comparison with a feedback linearizing control de-
sign. Again using z1 = γ − γref , feedback linearization
proceeds by defining References
[1] R.P. Brent. Algorithms for Minimization Without
ż1 = ϕ(α − α0 ) = z2 Derivatives. Prentice-Hall, 1973.
ż2 = ϕ0 (α − α0 )(q − z2 ) = z3 [2] G.E. Forsythe, M.A. Malcolm, and C.B. Moler.
ż3 = ϕ00 (α − α0 )(q − z2 )2 + ϕ0 (α − α0 )(u − z3 ) = ũ Computer Methods for Mathematical Computations.
Prentice-Hall, 1976.
 T
Selecting ũ = − k1 k2 k3 z1 z2 z3 clearly [3] A.J. Krener and A. Isidori. Linearization by out-
renders the closed loop system linear in the z coor- put injection and nonlinear observers. Systems & Con-
dinates. Solving for the angular acceleration u to be trol Letters, 3:47–52, June 1983.
produced, defined as in (15), we get
[4] M. Krstić, I. Kanellakopoulos, and P. Kokotović.
ũ − ϕ00 (α − α0 )(q − z2 )2 Nonlinear and Adaptive Control Design. John Wiley &
u = z3 + (24) Sons, 1995.
ϕ0 (α − α0 )
[5] M. Krstić and P.V. Kokotović. Lean backstepping
Two things are worth noting about this expression. design for a jet engine compressor model. In Proceedings
First, it depends not only ϕ, but also on its first and of the 4th IEEE Conference on Control Applications,
second derivatives. Recalling its definition from Equa- pages 1047–1052, 1995.
tion (7), this means that the lift force L must be well [6] S.H. Lane and R.F. Stengel. Flight control de-
known in order to accurately compute u. In the back- sign using non-linear inverse dynamics. Automatica,
stepping design, we relied on L only through its zero 24(4):471–483, 1988.
crossing and its sign switching property (8). Second, ϕ0
is in the denominator of the expression above, implying [7] T. Larsson. Linear quadratic design and µ-
that the control law has a singularity where ϕ0 = 0. analysis of a fighter stability augmentation system.
This occurs around the stall angle, where the lift force Master’s thesis, Linköpings universitet, 1997.
no longer increases with α, see Figure 2. [8] J-F. Magni, S. Bennani, and J. Terlouw, editors.
Robust Flight Control - A Design Challenge. Springer,
1997.
7 Conclusions [9] Jean-Jaques E. Slotine and Weiping Li. Applied
Nonlinear Contol. Prentice Hall, 1991.
In this paper we have applied backstepping to flight [10] B.L. Stevens and F.L. Lewis. Aircraft Control and
path angle control. Regarding the pitching moment as Simulation. John Wiley & Sons, 1992.

You might also like